New angles in nuclear magnetic resonance sample spinning

Gordon E. Brown, Jr. , Victor E. Henrich , William H. Casey , David L. Clark , Carrick .... Larsen , K. Schmidt‐Rohr , M. Hong , L. Frydman , G. C. ...
0 downloads 0 Views 2MB Size
VOLUME 25 NUMBER 5

MAY 1992 Registered in US.Patent and Trademark Office; Copyright 1992 by the American Chemical Society

New Angles in Nuclear Magnetic Resonance Sample Spinning E. WRENNWOOTEN,? KARLT. MUELLER,~ and ALEXANDER PINES* Materials Sciences Division, Lawrence Berkeley Laboratory, 1 Cyclotron Road, Berkeley, California 94720, and Department of Chemistry, University of California, Berkeley, California 94720 Received October 26, 1991 (Revised Manuscript Received December 5, 1991)

Introduction The resolving power and site specificity inherent in nuclear magnetic resonance (NMR) spectroscopy have helped make this technique one of the most versatile probes available to chemists. With NMR one hopes to assign clearly identifiable transitions, or resonances, to individual nuclei within a molecule and, from these resonances, to acquire a detailed understanding of local chemical and physical perturbations. Many different types of systems can be studied in thisway, solid as well as liquid, but for each a key consideration is resolution. How closely and how specifically any material can be studied by NMR depends largely on the possibility of obtaining well-separated peaks corresponding to distinct atomic sites. High-resolution NMR of the liquid state has become widespread partly because intrinsic line widths are very narrow, and also because commercial spectrometers able to realize this natural potential have been available for many years. The chief reason why NMR lines are so narrow for liquids is motional averaging, as brought about by rapid molecular reorientation. The principal spin interactions, among which are the chemical shift, spin-pin scalar (or J) coupling, electric quadrupole coupling, and magnetic dipole-dipole coupling, all transform as tensors under rotations, and thus their E. Wrenn Wooten Qrew up In Arkansas and recelved hIs B.S. In chemistry from the University of the South In 1986. He then attended the Universlly of Oxford as a Rhodeg scholar and recelved Ms D.PhlI. In 1989. He spent two years at W e l e y as an NIH postdoctoral felbw and has recently Jolned the biophysks and phannacy facultks at the Unhrersity of Michlgsn: Karl T. Mudkw, a native of Tonawanda, New Y&, recehral hIs B.S. In chemistry from the University of Rochester In 1985. After attending Cam bridge Unhnwslty as a ChurdvM scholar, he came to Berkeley as an NSF graduate fellow and recelved his Ph.D. In 1991. He is cwently an NSERC po&doctoml f e k w at the unlverslty of Brftish Columbia Department of Chem Istry. Alexander P h Is Professof of ChemIsby at the Unlverslty of Callfomla and Senior sclentlst at the lnmence Berkeley Laboratory. He le weMtnown for his contributkns to NMR spectroscopy.

magnitudes depend on molecular orientation.' Consider, for example, the familiar chemical shift interaction. This interaction, by which the nucleus is partially shielded from the external magnetic field owing to the response of the surrounding electrons, gives rise to the frequency dispemion seen in an NMR spectrum and can often be correlated with molecular topology and structure.2 The value of the chemical shift is determined by the position of each molecule or crystallite in the sample with respect to the external magnetic field. These positions would be f i e d in a rigid lattice, but in a liquid, where molecules are tumbling rapidly compared with the Larmor frequency, each molecule is able to sample all possible orientations on a short time scale. The spectral line that reaulta is a single peak reflecting the average of the chemical shift tensor over a sphere. This average quantity is independent of orientation, in much the same way as an s orbital has no angular dependence, and is called, accordingly, the isotropic chemical shift. Although most nuclei in solids typically are affected by chemical shift anisotropy (CSA), in some cases molecular tumbling is still sufficiently fast that narrow NMR lines are also obtained. It was this characteristic of adamantane that made it so desirable for the early high-resolution solid-state 13C cross-polarization experiment~.~Molecules of the newly discovered Cso *Towhom correapondence should be addressed. 'Present address: Biophysica Research Division, The University of Michigan, Ann Arbor, MI 48109. Pment address: Department of Chemistry, University of British Columbia, Vancouver, British Columbia, V6T 1Z1 Canada. (1) Mehring, M. Principles of High Resoh&~n NMR in Solids, 2nd ed.; Springer-Verlag: New York, 1983. (2) Silverstein, R. M.; Bassler, G. C.; M o d , T. C. Spectrometric Identification of Organic Compounds;John Wdey and Sons,Inc.: New York, 1981. (3) Pines, A.; Gibby, M. G.; Waugh, J. S. J. Chem. Phys. 1972,56, 1776-1777.

*

0 1992 American Chemical Society

Wooten et al.

210 Acc. Chem. Res., Vol. 25, No. 5, 1992

!

the electrons surrounding it. The resulting orientation-dependent splitting of the spectral line provides information about bond order and site sy”etrye7 Figure 2 shows how this interaction is averaged under tetrahedral jumps, which give rise to a tetrahedral pseudorotation of individual water molecules. Since most solids at room temperature exhibit broad spectral lines, much effort over the last 30 years has gone into developing ways to average the anisotropic broadening artificially, either by manipulating the spins or by manipulating the sample. One important class of the latter approach involves imposing macroscopic motion on the solid sample in an attempt to mimic the microscopic motion of the molecules in liquids. Appropriate motion of the sample container can yield narrowing comparable to that found in solution. In what follows, we discuss some recent developments in sample reorientation techniques which enable highresolution spectra to be obtained from a wide range of nuclei in the solid state.

295 K

A -

h

I\

400

200

123 K

0

- 200

ppm (TMS)

Figure 1. ’% NMR spectra of solid CW(buckminsterfullerene). At low temperatures the spectra are broad, refleeting the chemical shift anisotropy of the carbon-13sites in the rigid molecules. At higher temperatures the molecules reorient rapidly, averaging the anisotropy and leaving a sharp resonance line at the isotropic chemical shift. Reprinted with permission from ref 4. Copyright 1991 American Chemical Society.

buckminsterfullerene, with the shape of a truncated icosahedron, also tumble rapidly at room temperature and thus produce narrow resonances even in the solid state4J’(Figure 1). In most solids, however, molecular reorientation is not sufficiently fast to average the spin interactions, and broad lines result. One can envision the broad solid-state line as being composed of a series of narrow lines, each coming from a different crystallite orientation. Indeed, as shown in Figure 1,such is the case for CWat lower temperatures. The regime between fast tumbling and slow tumbling can be exploited to distinguish different types of molecular dynamics. For example, Wittebort et al.6 used line-shape simulations of proton and deuterium spectra taken over a range of temperatures to show that reorientation of the hydrogen bonds in polycrystalline hexagonal ice occurs via tetrahedral jumps rather than continuous rotational diffusion. For I > 1/2 nuclei, such as deuterium (I= l), an electric quadrupole interaction arises from the coupling of a nonspherical nuclear charge distribution with the electric field gradient of (4) Yannoni, C. S.;Johnson, R. D.; Meijer, G.; Bethune, D. S.; Salem, J. R. J.Phys. Chem. 1991,95,9-10. (5) Tycko, R.; Haddon, R. C.; Dabbagh, G.; Glarum, S. H.; Douglass, D. C.; Mujsce, A. M. J. Phys. Chem. 1991,95,518-520. (6) Wittebort, R. J.; Usha, M. J.; Ruben, D. J.; Wemmer, D. E.; Pines, A. J.Am. Chem. SOC.1988,110, 5668-5671.

Averaging First-Order Interactions: Magic-Angle Spinning Approximately 100 of the various atomic nuclei possess a nonzero spin angular momentum, characterized by a spin quantum number I. Of these, over half are quadrupolar, i.e., they have I > 1/2. When a spin I is placed in an external magnetic field, the spin states are dispersed over 21 + 1 energy levels (each characterized by a quantum number m) due to the Zeeman interaction. In addition, the spins are subject to a wide range of interactions among themselves and with their environment, such as chemical shift, J coupling, dipoledipole coupling, and quadrupole coupling. Consider now the case of a sample contained in a rotor spinning rapidly about an axis inclined at an angle with respect to the external magnetic field. A calculation of the average resonance frequencies of the nuclei from first-order static perturbation theory or from coherent averaging the^@^ gives the following resulta for the first-order chemical-shift (CS), scalar-coupling(4, dipolar-coupling (DD), and quadrupolar-coupling (Q) interactions: =

+ w p p 2e)( ~ ~ ~

(1)

where X = CS, J, DD, or Q. 0 is the angle between the sample-spinner axis and the external magnetic field. The chemical-shift and scalar-coupling interactions consist of an isotropic (orientation-independent) part and an anisotropic (orientation-dependent) part, while the dipolar and first-order quadrupole interactions contain only an anisotropic part (w&ib,iso= 0, w @ L= 0). In all four cases the anisotropic contribution is scaled by P2(cose), the second-order Legendre polynomial of cos 8. A useful way to picture the spatial distribution of these interactions is to think of the isotropic parta as s orbitals whose radii determine the strength of the interaction, and to envision the anisotropic parts as dZz orbitals (Figure 3). While there is, of course, no spatial (7) Kirkpatrick, R. J. MAS NMR Spectroscopy of Minerals and Glasses. In Spectroscopic Methods in Mineralogy and Geology; Hawthorne, F. C., Ed.; Reviews in Mineralogy; MineralogicalSociety of America: Washington, DC, 1988, Vol. 18. (8) Haeberlen, U. Advances in Magnetic Resonance, Supplement I ; Academic Press: New York, 1976. (9) Maricq, M. M.; Waugh, J. S. J. Chem. Phys. 1979,70,330&3316.

Acc. Chem. Res., Vol. 25, No. 5,1992 211

New Angles in NMR Sample Spinning

Simulation (b)

Experimental (a)

267K

262K

258K

z49K

lL

A

I = 1.8%

---2.3%

A